Sei sulla pagina 1di 9

Journal of Hazardous Materials 179 (2010) 564572

Contents lists available at ScienceDirect


Journal of Hazardous Materials
j our nal homepage: www. el sevi er . com/ l ocat e/ j hazmat
Application potential of grapefruit peel as dye sorbent: Kinetics, equilibrium and
mechanism of crystal violet adsorption
Asma Saeed
a,
, Mehwish Sharif
b
, Muhammad Iqbal
a
a
Environmental Biotechnology Group, Biotechnology and Food Research Centre, PCSIR Laboratories Complex, Ferozepur Road, Lahore 54600, Pakistan
b
School of Biological Sciences, University of the Punjab, Lahore 54590, Pakistan
a r t i c l e i n f o
Article history:
Received 16 November 2009
Received in revised form 20 February 2010
Accepted 10 March 2010
Available online 16 March 2010
Keywords:
Dye sorption
Grape fruit peel
Crystal violet
Fixed-bed column
a b s t r a c t
This study reports the sorption of crystal violet (CV) dye by grapefruit peel (GFP), which has application
potential in the remediation of dye-contaminated wastewaters using a solid waste generated by the
citrus fruit juice industry. Batch adsorption of CV was conducted to evaluate the effect of initial pH,
contact time, temperature, initial dye concentration, GFP adsorbent dose, and removal of the adsorbate
CV dye from aqueous solution to understand the mechanism of sorption involved. Sorption equilibrium
reached rapidly with 96% CV removal in 60min. Fit of the sorption experimental data was tested on
the pseudo-rst and pseudo-second-order kinetics mathematical equations, which was noted to follow
the pseudo-second-order kinetics better, with coefcient of correlation 0.992. The equilibrium process
was well described by the Langmuir isothermmodel, with maximumsorption capacity of 254.16mgg
1
.
The GFP was regenerated using 1M NaOH, with up to 98.25% recovery of CV and could be reused as a
dye sorbent in repeated cycles. GFP was also shown to be highly effective in removing CV from aqueous
solution in continuous-ow xed-bed column reactors. The study shows that GFP has the potential of
application as an efcient sorbent for the removal of CV from aqueous solutions.
2010 Elsevier B.V. All rights reserved.
1. Introduction
The global problem of clean water shortage has been exac-
erbated by its pollution caused through discharge of untreated
industrial efuents. The problem is further aggravated in devel-
oping countries owing to uncontrolled population growth, and the
use of outdated practices and technologies that consume large vol-
umes of water in agricultural and industrial operations [1]. Several
industrial processes utilize toxic chemicals for the manufacture
of nished products, the unused parts of which escape into the
environment as industrial waste-wash [2]. Environmental scien-
tists worldwide are thus faced withthe daunting task of developing
cost-effective andefcient technologies for their treatment for mis-
cellaneous human activities [3]. Discharge of industrial efuents
containing hazardous contaminants, such as phenolics, toxic met-
als and dyes, even at low concentrations, are a cause of negative
impact on the environment [4,5]. Out of the various pollutants,
dyes rank among the most notorious organic contaminants that
are discharged into the environment during textile, leather, paints,
paper and dye manufacturing [5,6]. These are synthetic, water sol-
uble and dispersible organic compounds, which may adversely

Corresponding author. Tel.: +92 42 99230688; fax: +92 42 99230705.


E-mail address: asmadr@wol.net.pk (A. Saeed).
affect aquatic ora andfauna by reducing light penetrationthrough
water surface and their toxic effects, thus causing severe dam-
age to aquatic biota [7]. Dyes are extremely stable due to their
complex aromatic molecular structure, and are thus difcult to
biodegrade [8,9]. From the aspect of environmental safety, the
removal of synthetic dyes fromthe industrial waste-washis of seri-
ous concern, since some dyes and their degradation products are
toxic and carcinogenic [4,10]. Various conventional technologies,
such as chemical oxidation and reduction, physical precipitation
and occulation, photolysis, adsorption, electrochemical treat-
ment, reverse osmosis, bioaccumulation, and/or biodegradation,
have beeninvestigatedfor their removal fromwastewaters [11,12].
Most of these methods, nevertheless, pose techno-economical lim-
itations for eld-scale applications [13].
Crystal violet dye, a member of the triphenylmethane group, is
extensively used in animal and veterinary medicine as a biolog-
ical stain, for identifying the bloody ngerprints being a protein
dye, and in various commercial textile operations [14]. It is car-
cinogenic and has been classied as a recalcitrant molecule since it
is poorly metabolized by microbes, is non-biodegradable, and can
persist inavarietyof environments [11]. Its removal fromwastewa-
ters before their discharge is, therefore, essential for environmental
safety. Adsorption has been reported to be efcient and econom-
ical for the treatment of wastewaters containing dyes, pigments
and other colourants [13]. Granular activated carbon has been used
successfully, but is cost-prohibitive [15]. This has led to search for
0304-3894/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jhazmat.2010.03.041
A. Saeed et al. / Journal of Hazardous Materials 179 (2010) 564572 565
cheaper adsorbent substitutes derived from agricultural and wood
wastes [4]. As a result, activated carbon made from male owers
of coconut [14], rice husk [16] and ame tree (Delonix regia) pods
[4], and acid-treated neem (Azadirachta indica) sawdust [5] and
almond skin [9], have been studied for the sorption of crystal vio-
let. These adsorbents either involve cost-input pretreatments or are
not abundantly available as wastes. The present report attempts to
eliminate such preparatory steps and limitations of availability by
using raw grapefruit peel, a waste generated in abundance during
industrial processing of the fruit.
Grapefruit is cultivated in all tropical and subtropical regions
of the world, with approximately 4 million metric tons annual
production [17]. Thus, a lot of grapefruit peel (GFP) is available
throughout theworld. GFPcontains several water solubleandinsol-
uble monomers and polymers [18,19]. The water soluble fraction
contains glucose, fructose, sucrose and some xylose, while pectin,
cellulose, hemicellulose andligninconstitute between50%and70%
of the insoluble fraction. These polymers are rich in carboxyl and
hydroxyl functional groups, whichmaybindcationic dyemolecules
inaqueous solution[20]. The present work was undertakento eval-
uate the application potential of the pectin- and cellulose-rich GFP
as an inexpensive and environment-friendly adsorbent for treat-
ing wastewaters containing the cationic crystal violet (CV) dye.
The study is further unique as there is no existing report for the
removal of any dye by GFP. Parameters investigated for sorption
of the dye included the effect of temperature, pH, biomass dose,
and dye concentration. FTIR analysis was done to identify the func-
tional groups present inthe grapefruit peel, anddye desorptionwas
done to determine reuseability of the sorbent peel for a number of
repeated applications. An attempt has been made to interpret the
data, based on ion exchange mechanism, and for nding a t on
various kinetics and isotherms equations.
2. Material and methods
2.1. Sorbent material
Grapefruit peel was collected from the local fresh fruit juice
vending outlets. The GFP was sorted by removing leaves, twigs and
other debris immediately after collection. After thorough washing
with tapwater, GFP was washed with double distilled water, oven
driedat 702

C, ground, andsieved. PowderedGFP of particle size


0.851.0mm was used for the sorption studies.
2.2. Crystal violet stock solution
Crystal violet, a basic dye, C.I. 42555,
max
=586nm, molecu-
lar formula C
25
H
30
N
3
Cl, also known as hexamethyl pararosaniline
chloride, was purchased from Fluka, Darmstadt, Germany and its
chemical structure is shown in Fig. 1. Stock solution of 1000mgL
1
crystal violet was prepared by dissolving 1.0gL
1
of the dye in
Fig. 1. Chemical structure of crystal violet.
ultrapure deionized water (18Mcm). The dye solution pH was
adjusted using 0.1N HCl or 0.1N NaOH. Fresh dilutions of the
desired dye concentrations were made at the start of each experi-
ment.
2.3. Sorption procedure for equilibrium and kinetics studies
The maximum sorption capacity was determined by con-
tacting 100mL crystal violet solution of known concentrations
(5600mgL
1
, pH 6.0) in 250-mL Erlenmeyer asks. The asks,
tightly stoppered, were shaken on orbital shaker at 100rpm at
ambient temperature (302

C) for 60min to ensure establish-


ment of equilibrium. The GFP was separated by centrifugation at
5000rpm for 5min. Residual concentration of crystal violet was
determined spectrophotometrically (Shimadzu spectrophotome-
ter UVVis 1700, Japan) at 585nm. Dilutions were made when
absorbance exceeded 1.5. The effect of initial pH was determined
by varying dye solution pH in the range 210. Biomass concentra-
tion was optimized by varying GFP dose (0.13.0gL
1
). Dye-free
solution (double distilled water) was used as control. The effect
of temperature on sorption was determined at 20, 30 and 45

C.
The amount of crystal violet sorbed onto GFP, q
e
(mgg
1
), was
calculated using the equation:
q
e
=
(C
i
C
e
)V
W
(1)
where q
e
is the crystal violet uptake (mgdye g
1
sorbent); C
i
and
C
e
are, respectively, the initial and equilibrium liquid-phase con-
centrations of crystal violet (mgL
1
); V the volume of crystal violet
solution (L); and W the weight of GFP (g).
For determining the optimumcontact time for sorption equilib-
rium and kinetics of sorption, the sorbatesorbent were contacted
for various time intervals (10240min). The residual dye concen-
tration was determined and the amount of crystal violet sorbed, at
time t was calculated using the equation:
q
t
=
(C
i
C
e
)V
W
(2)
where q
t
is the crystal violet uptake at time t (mgdye g
1
sorbent).
2.4. Desorption and reuse of regenerated sorbents
The CV-loaded biomass of GFP, which was initially exposed to
100mgL
1
of CV at pH 6.0 and 30

C, was contacted with 50mL of


1MNaOHas the dye-desorbing agent for 60min on a rotary shaker
at 100rpm. The GFP biomass was removed by centrifugation and
the quantity of crystal violet dye recovered was determined. After
desorption, the GFP was washedseveral times withdouble distilled
water and the regenerated GFP was reused for the next adsorption
cycle following the same procedure as employed in the adsorption
equilibrium experiments.
2.5. FTIR analysis
FTIR spectrophotometer (ThermoNicolet IR-100 spectrometer,
ThermoNicole Corporation, Madison, USA) was used to identify
the chemical functional groups present on the native GFP. IR
absorbance data were obtained for wavenumbers in the range
of 4004000cm
1
and analyzed using Encompass

software pro-
vided by the FTIR spectrophotometer manufacturer. For obtaining
FTIR spectra, ball-milled GFP biomass was mixed with KBr (ratio
1:100), compacted into a tablet form using a bench press, and the
material was run for obtaining FTIR spectra.
566 A. Saeed et al. / Journal of Hazardous Materials 179 (2010) 564572
2.6. Column studies
A glass column, 1.7cm in diameter 25cm in height, was
packed with 140.56 GFP having 20cm bed-height. CV solu-
tion of known concentration (50mgL
1
or 100mgL
1
, pH 6.0)
was then pumped upwards through the column at a ow rate of
2.5mL min
1
. Samples were collected at regular intervals from the
efuent to measure residual CV concentrations. As the bed was
saturated, the CV loading was terminated. The total amount of CV
uptake by the GFP-packed column was calculated according to the
mass balance of dye expressed as:
q =
V(C
i
C
eq
)
M
(3)
where V is the volume of dye solution (L) passed through the col-
umn; C
i
the concentration of dye in the inlet solution; C
eq
the dye
concentrationof solutionat outlet of thecolumn; andMtheamount
of GFP (g) packed in column.
2.7. Reproducibility and data analysis
Unless indicated otherwise, the data reported are the mean val-
ues of three separate experiments. The amount of dye adsorbed
per unit biomass (mgdye g
1
GFP) was determined as shown in
Eq. (1). Langmuir and Freundlich models [21,22] were used for the
evaluationof sorptiondata. Langmuir isothermassumes monolayer
adsorption, which is presented by the equation:
q
eq
=
q
max
bC
eq
(1 +bC
eq
)
(4)
where q
eq
and q
max
are the equilibrium and maximum dye uptake
capacities (mgg
1
sorbent), respectively; C
eq
the equilibrium dye
concentration (mgL
1
); and b the Langmuir equilibrium constant
(L mg
1
).
Freundlich model is presented as:
q
eq
= K
F
C
1/n
eq
(5)
where K
F
and n are the Freundlich constants characteristic of the
system.
In order to examine the controlling mechanism of the sorption
process, such as mass transfer and chemical reaction, the pseudo-
rst-order, thepseudo-second-order, andintraparticulatediffusion
models wereusedtotest thet of experimental dataof dyesorption
by GFP on the kinetics equations proposed by various authors. The
pseudo-rst-order rate equation of Lagergren [23] is presented as:
ln(q
eq
q
t
) = lnq
eq
k
1ad
t (6)
where k
1ad
(min
1
) is the pseudo-rst-order reactionrate constant.
The pseudo-rst-order considers the rate of occupation of sorp-
tion sites to be proportional to the number of unoccupied sites.
A straight line of ln(q
eq
q
t
) versus t indicates the application of
pseudo-rst-order kinetics model. Ina true pseudo-rst-order pro-
cess, lnq
eq
should be equal to the intercept of a plot of ln(q
eq
q
t
)
against t. Thepseudo-second-order equation[24], another equation
used for kinetics analysis, which is based on the sorption equilib-
rium capacity, may be expressed in the following form:
t
q
t
=
1
k
2ad
q
2
eq
+
1
q
eq
t (7)
where k
2ad
is the pseudo-second-order rate equilibrium constant
(gmg
1
min
1
).
A plot of t/q
t
against t should give a linear relationship for the
applicability of the pseudo-second-order kinetics model.
Fig. 2. Effect of initial pH on the sorption of crystal violet by grapefruit peel waste.
3. Results and discussion
3.1. Effect of pH on dye sorption
The pH of an aqueous solution is an important factor in dye
adsorption, as it affects the surface charge of the sorbent material
and the degree of ionization of the dye molecule [4]. pH has been
related with changes in the structural stability and colour intensity
of the dye molecule [6]. The removal of CV by GFP over a range
of 210 was noted to increase with the increase in pH of the dye
solution, appreciably up to pH 6.0 (Fig. 2). A further increase in
dye sorption between pH 6.0 and 10 was insignicant. Since the
optimum pH for dye biosorption by GFP was found to be 6.0, this
pH was used for further studies. As the pH increases, the charge
density of the dye solutiondecreases, sothat electrostatic repulsion
between the positively charged dye molecule and the surface of
the adsorbent is lowered [6], which results in an increase in the
sorption of the dye. The minimumremoval of CV was thus found at
pH 2, which is probably due to the H
+
ions competing favourably
with the cationic groups of the dye molecule for sorption sites on
GFP biomass [6]. Similar observations have been reported by other
authors [4,6]. The insignicant higher removal of the dye, observed
at pH between 6 and 10, was noted to be accompanied by a drop
in the original pH of the dye solution at sorption equilibrium. This
indicates a cationic exchange process between the dye molecule
and the GFP biomass with the release of protons (H
+
), as suggested
in an earlier study [25].
3.2. Effect of contact time
The rate of sorption of CV by GFP was determined by contacting
25mgL
1
of the CVsolution(pH6.0) with1.0gL
1
GFP for different
intervals of time. The equilibriumwas reached within 60min of the
sorbatesorbent contact (Fig. 3). The uptake of CV, as a function of
time, was noted to occur in two phases. The rst phase involved a
rapid uptake of dye during the rst 20min of the sorbatesorbent
contact, which was followed by a slowphase of dye removal spread
over a signicantly longer period of time (>60min) until the equi-
librium was reached. The two-stage sorption, the rst rapid and
quantitatively predominant and the second slower and quantita-
tively insignicant, has beenextensively reportedinliterature [26].
The rapid stage, furthermore, may last for several minutes to a few
hours, while the slowone continues for several hours to a day [27].
The rapid phase is probably due to the abundant availability of
active sites on the sorbent, whereas with the gradual occupancy
of these sites the sorption process becomes less efcient during
the slower phase [28].
A. Saeed et al. / Journal of Hazardous Materials 179 (2010) 564572 567
Fig. 3. Biosorption of crystal violet by grapefruit waste as a function of time.
3.3. Effect of temperature on dye adsorption
Effect of temperature on the adsorption of CV by GFP at equilib-
rium was investigated at three different temperatures, viz., 20

C,
30

C and 45

C at the initial CVconcentration of 25 and 100mgL


1
,
pH 6.0 and the adsorbent dose of 1.0gL
1
(Fig. 4). The adsorp-
tion of CV was noted to enhance with the increase in temperature
indicating that higher temperature favoured the removal of dye by
sorption onto GFP. The observation is in agreement with the earlier
report on the adsorption of CV by activated carbon from rice husk
[16]. The enhanced adsorption at higher temperatures was sug-
gestedtobe due toincrease inthe availability of active surface sites,
increased porosity, and in the total pore volume of the adsorbent.
These authors further suggested that the enhancement in adsorp-
tion may also be a result of an increase in the mobility of the dye
molecule with an increase in their kinetic energy, enhanced rate of
intraparticulate diffusion of the sorbate dye, and due to decrease
in thickness of the boundary layer surrounding the sorbate dye
at higher temperature so that the mass transfer resistance of the
sorbate in the boundary layer decreases. The observations on the
adsorption of CV on GFP in the present study further suggest that
it was an endothermic process and may involve not only physical
but also chemical sorption, as suggested in an earlier study on the
adsorption of CV by activated carbon derived from male owers of
coconut tree [14].
3.4. Effect of the solid sorbent to the liquid sorbate ratio on CV
adsorption
One of the parameters that strongly affect sorption capacity is
the quantity of the contacting sorbent in the liquid phase [26].
Fig. 4. Effect of temperature on the equilibriumsorption capacity of grapefruit peel.
Fig. 5. Effect of grapefruit peel waste dosage on the sorption of crystal violet.
The inuence of the solid sorbent to the liquid sorbate ratio on
the sorption capacity of GFP was investigated at the constant ini-
tial concentration of CV 25mgL
1
(liquid phase), whereas the GFP
mass (solid phase) was varied between 0.1gL
1
and 3.0gL
1
. The
results obtained during the study show that increase in the solid
phase mass from0.1gL
1
to 1.0gL
1
resulted in a rapid increase in
the uptake of CV (Fig. 5). Further increase in the solid phase mass
from 1.0gL
1
to 3.0gL
1
, however, did not result in any apprecia-
ble increase in the sorption capacity of GFP. Therefore, the solid
phase mass of 1.0gL
1
was selected for further studies. It may
be concluded from these observations that at lower GFP dosage of
<1.0gL
1
, the dye molecules were competing for sorption at limit-
ing sorption sites. Therefore, as the quantity of GFP was increased
from 0.1gL
1
to 1.0gL
1
, the availability of sorption sites eased
resulting in greater percentage removal of the dye. The insigni-
cant increase in the uptake of dye at the GFP dosage higher than
1.0gL
1
may be attributed to the presence of excess/surplus dye-
binding sites on GFP than the available dye molecules present in
the solution at the xed concentration of 25mgL
1
. These obser-
vations are in agreement with those reported previously by other
researchers for the sorptionof dyes bydifferent biological materials
[7,13].
3.5. Effect of initial dye concentration on CV sorption
Dye adsorption capacity of GFP, as a function of initial con-
centration of dye within the aqueous solution is shown in Fig. 6.
The initial concentration of dye was changed in the range of
10600mgL
1
. The amount of dye adsorbed per unit mass of the
adsorbent increasedwithincrease inthe initial concentrationof CV.
In order to reach the plateau values, which represent saturation
Fig. 6. Effect of initial dye concentration on crystal violet biosorption by grapefruit
peel waste.
568 A. Saeed et al. / Journal of Hazardous Materials 179 (2010) 564572
of the active sites on the sorbent, that is, the maximum adsorp-
tion capacity of the adsorbent, the initial concentration of CV was
increased up to 600mgL
1
. As may be noted from Fig. 6, the maxi-
mumamount of dye adsorbed on GFP was 254mgg
1
. The value is
muchhigher thanthe other biowaste materials previously reported
as adsorbents for the removal of dyes, such as, 85.47mgg
1
by
skin almond waste [9], 60.42mgg
1
by activated carbon frommale
owers of coconut tree [14], and 64.87mgg
1
by activated carbon
prepared from rice husk [16].
3.6. Adsorption isotherms modeling
Adsorptiondata are most commonly representedby the equilib-
rium isotherm value, which is a plot of the quantity of the sorbate
removed per unit sorbent (q
eq
) as the solid phase concentration
of the sorbent against the concentration of the sorbate in the liq-
uid phase (C
eq
). The equilibrium isotherm value is of fundamental
importance for the design and optimization of the adsorption sys-
tem for the removal of a dye from an aqueous solution. Therefore,
it is necessary to establish the most appropriate correlation for
the equilibrium curve. Several isotherm models have been used
to predict validity of the experimental data. In the present study,
two of the most commonly used models, namely, the Langmuir
and Freundlich isotherms, were used to describe the adsorption
equilibrium.
TheLangmuir isothermis basedontheassumptionof monolayer
adsorptionona structurally homogeneous adsorbent, where all the
adsorptionsites are identical andenergetically equivalent, wherein
the adsorption occurs at specic homogeneous sites within the
adsorbent, and once a dye molecule occupies a site no further
adsorption can take place at that site [6]. The intermolecular forces
decrease rapidly with distance, and consequently the existence of
monolayer coverage of the adsorbate at the outer surface of the
adsorbent can be predicted [21]. The Freundlich isotherm model
is an empirical expression that encompasses the heterogeneity of
the surface and the exponential distribution of sites and their ener-
gies. Acomparisonof the Langmuir (Fig. 7a) andFreundlich(Fig. 7b)
adsorption isotherms shows that the sorption characteristics of CV
dye onto GFP followed more closely the Langmuir isotherm equa-
tion (Eq. (3)) than the Freundlich isotherm equation (Eq. (4)). This
observationis further supportedby the closer tounity value of their
respective correlation coefcients (r
2
), which is a measure of how
well the predicted values from a forecast model match with the
experimental data [29]. The adsorptionparameters of CVonGFP for
Langmuir andFreundlichconstants aregiveninTable1. Ther
2
value
inrespect of sorptionof CVfor Langmuir isothermmodel was noted
to be 0.994, which for Freundlich isothermmodel was 0.974. It may
be concluded fromthese observations that the adsorption of CV by
GFP was better dened by the Langmuir than by the Freundlich
equation thus indicating that the adsorption of CV onto GFP is a
chemically equilibrated and saturated mechanism. The maximum
CV binding capacity of GFP was found to be 249.68mgg
1
and was
consistent with the experimental data (Table 1). This adsorption
capacity of GFP for CV was also found to be signicantly higher in
comparison with some other recently reported adsorbents, such as
85.47mgg
1
by skin almond waste [30], 64.87mgg
1
by activated
rice husk [16], and 113mgg
1
by activated sludge [31].
Fig. 7. The linearized (a) Langmuir and (b) Freundlich adsorption isotherms for the
sorption of crystal violet by grapefruit peel.
Table 2
Characteristics of the Langmuir adsorption
isotherms.
Separation factor (RL) Type of isotherms
RL >1 Unfavourable
RL =1 Linear
0>RL <1 Favourable
RL =0 Irreversible
The Langmuir parameters can also be used to predict afnity
between the sorbate and the sorbent using the dimensionless sep-
aration factor (R
L
), which has been dened as below [32]:
R
L
=
1
1 +bC
i
(8)
The value of R
L
canbe used to predict whether a sorptionsystem
is favourable or unfavourable in accordance with the criteria
shown in Table 2. The value of R
L
for the sorption of CV onto GFP
is shown in Fig. 8, which indicates that sorption of CV on GFP was
favourable.
3.7. Kinetics modeling
To examine the controlling mechanism of the adsorption
process, such as mass transfer and chemical reaction, the pseudo-
rst-order and the pseudo-second-order kinetics models were
used to test the experimental data of dye adsorption by GFP. The
Table 1
The Langmuir and Freundlich isotherms model constants, and their respective correlation coefcients (r
2
) for the sorption of crystal violet fromaqueous solution by grapefruit
peel.
Experimental (mgg
1
) Langmuir Freundlich
qmax (mgg
1
) b (L mg
1
) r
2
KF 1/n r
2
254.166.86 249.68 0.131 0.994 3.14 0.446 0.974
A. Saeed et al. / Journal of Hazardous Materials 179 (2010) 564572 569
Table 3
Theoretically determined constants of the pseudo-rst and the pseudo-second-order reaction kinetics based on the sorption of crystal violet from 25mgL
1
solutions, pH
6.0, by 1.0gL
1
grapefruit peel during shake ask sorbentsorbate contact at 100rpm for 240min.
Experimental qeq (mgg
1
) Pseudo-rst-order constants Pseudo-second-order constants
qeq (mgg
1
) k
1
(min
1
) r
2
qeq (mgg
1
) k
2
(gmg
1
min
1
) r
2
24.05 8.84 0.031 0.838 24.31 0.005 0.992
Fig. 8. Value of separation factor RL for sorption of crystal violet by grapefruit peel.
experimental data for the removal of CV when analyzed on the
pseudo-rst-order equation (Eq. (5)) did not result in a perfect
straight line when a plot was drawn between ln(q
eq
q
t
) and t
indicating a non-t on the model (Fig. 9a). A good linear plot of
t/q
t
against t for the pseudo-second-order kinetics model (Eq. (6))
shows a t on the model (Fig. 9b). Kinetics parameters for the
sorption of CV by GFP, as calculated from the linear plots of the
Fig. 9. Linearized (a) pseudo-rst-order and (b) pseudo-second-order plots for the
sorption of crystal violet by grapefruit peel waste.
pseudo-rst-order (Fig. 9a) and the pseudo-second-order (Fig. 9b)
kinetics models are presented in Table 3. The low correlation coef-
cient value (r
2
=0.838), as obtained for the pseudo-rst-order
model, indicates that sorption of CV did not follow the pseudo-
rst-order reaction. The insufciency of the pseudo-rst-order
model to t the kinetics data could possibly be due to the limi-
tations of the boundary layer controlling the sorption process. The
experimental data were observed to t well the pseudo-second-
order equation. Thehighcorrelationcoefcient value(r
2
=0.992), as
obtained for the linear plot of t/q
t
against t for the pseudo-second-
order equation, was observed to be close to 1. This suggests that
the process of sorption kinetics of CV by GFP follows the pseudo-
second-order equation and the process controlling the rate may be
controlled by chemical sorption involving valence forces through
sharing or exchange of electrons between sorbent and sorbate
[24].
3.8. Proposed sorption mechanism
The pattern of adsorption onto plant materials is attributable
to the active groups and bonds present on them [33]. For the
elucidation of these active site, FTIR spectrophotometry was per-
formed. Peaks appearing in the FTIR spectrum of GFP (Fig. 10)
were assigned to various groups and bonds in accordance with
their respective wavenumbers (cm
1
) as reported in literature.
The region between 2600cm
1
and 3600cm
1
shows two major
band stretches. Abroad and strong band stretch was observed from
3000cm
1
to 3600cm
1
, indicating the presence of free or hydro-
gen bonded OH groups (alcohols, phenols and carboxylic acids)
as in pectin, cellulose and lignin on the surface of the adsorbent
[34]. The OH stretching vibrations occur within a broad range
of frequencies indicating the presence of free hydroxyl groups
and bonded OH bands of carboxylic acids. The light stretch at
2909.64cm
1
showed the stretching of symmetric or asymmet-
ric CH vibration of aliphatic acids [35]. The peak observed at
1730cm
1
is the stretching vibration of C O bond due to non-
ionic carboxyl groups (COOH, COOCH
3
) and may be assigned
to carboxylic acids or their esters [36]. Asymmetric and sym-
Fig. 10. FTIR spectra of native grapefruit peel waste.
570 A. Saeed et al. / Journal of Hazardous Materials 179 (2010) 564572
Fig. 11. Proposed ion exchange mechanism between a proton of grapefruit peel waste and crystal violet dye.
metric stretching vibrations of ionic carboxylic groups (COO

),
respectively, appeared at 1641cm
1
and 1415cm
1
. The peaks at
1384cm
1
may be assigned to symmetric stretching of COO

of
pectin [36], and aliphatic acids group vibration at 1233cm
1
to
deformation vibration of C O and stretching formation of OH of
carboxylic acids andphenols [37], andat 1050cm
1
canbeassigned
to the stretching vibration of COH of alcoholic groups and car-
boxylic acids [37]. It is well indicated from the FTIR spectrum of
GFP that carboxyl and hydroxyl groups were present in abundance.
The sorption of CV on the GFP biomass may likely be due to the
electrostatic attraction between these groups and the cationic dye
molecule (CV
+
). At pH above 4, the carboxylic groups are depro-
tonated and negatively charged carboxylate ligands (COO

) bind
the positively charged CV molecules. This conrms that the sorp-
tion of CV by GFP was an ion exchange mechanism between the
negatively charged groups present in the cell wall of GFP and the
cationic dye molecule. The proposed mechanism of sorption of CV
by GFP is shown in Fig. 11.
3.9. Desorption of CV from GFP
Recovery of CV from exhausted biomass of GFP and regener-
ation of GFP was tested by using aqueous solution of 1M NaOH
[37]. For the purpose, 1g of CV-loaded GFP was contacted with
50mL of 1M NaOH and shaken at 100rpm on a rotary shaker at
30

C. The kinetics of desorption of CV from the CV-loaded GFP is


shown in Fig. 12. It may be noted that CV desorbed very rapidly,
with the maximum elution achieved in the rst 30min amount-
ing to 82% desorption. The desorption rate after this initial fast
phase, however, was observed to slow down signicantly until
it reached a plateau after 60min, indicating equilibrium of the
system. The desorption efciency was 98.25%. The regenerated
GFP biomass was again used for removal of CV and was found to
have the efciency of 87% and 81% in the second and the third
cycles, respectively. Similar observations for the desorption of CV
fromsurfactant-modied alumina using NaOHhave been reported
[38]. The successful recovery of CV at >98% using 1M NaOH from
the CV-loaded GFP and the high sorption q
max
of 254.16mgg
1
of GFP, coupled with the use of a no cost industrial waste as an
adsorbent, indicate that the material reported in the present study
has a good potential for the removal of the dye from wastewa-
ters.
3.10. CV removal by GFP from made-up wastewater in xed-bed
continuous-ow column
Thepotential shownbyGFPtoremoveCVfromaqueous solution
in shake ask studies was further tested in a xed-bed continuous-
ow column using made-up wastewater. For this purpose, CV
was added to tapwater at the known concentrations of 50mgL
1
and 100mgL
1
and passed through the separate columns packed
with 140.56g of GFP. The breakthrough biosorption curves are
presented in Fig. 13. The dye loading curve showed an excel-
lent clear zone (100% removal) before the breakthrough point.
The results presented in Fig. 13 show that the column bioreactor
packed with 140.56g GFP had the capacity to treat 61L and 29L
dye solution before reaching the breakthrough point for 50mgL
1
Fig. 12. Desorption kinetics of crystal violet from GFP.
A. Saeed et al. / Journal of Hazardous Materials 179 (2010) 564572 571
Fig. 13. Breakthroughcurves of biosorptionof crystal violet bygrapefruit peel waste
in xed-bed continuous-ow column.
and 100mgL
1
CV, respectively, and the saturation points were
achieved at 83L and 45L (Fig. 13). The total biosorption capac-
ity of the packed column bioreactor was calculated by integrating
the breakthrough curves between the breakthrough and satura-
tion points. The breakthrough point of the GFP-packed column is
the point at which CV concentration in the efuent (outlet dye
concentration) reached at concentration above zero whereas the
saturation point is the time at which CV concentration in the efu-
ent become equal to inlet dye concentration. The sorption capacity
of GFP for 50mgL
1
and 100mgL
1
CV, respectively, was calcu-
lated to be 259.91mgg
1
and 266.15mgg
1
, which is even better
than the maximum value of 254.16mgg
1
obtained during batch-
scale studies.
4. Conclusions
Studies suggest that GFP can be effectively used as a cost-
effective adsorbent for removal of CV fromaqueous solution. Batch
adsorption studies show that removal is dependent upon process
parameters likepH, sorbateandsorbent concentrations andcontact
time. Sorption followed pseudo-second-order kinetics equation.
The experimental equilibrium sorption data obtained from batch
studies at optimized conditions t well to Langmuir adsorption
isotherm equation, indicating monolayer adsorption. The dimen-
sionless parameter R
L
has also been calculated using the Langmuir
constant b. The values of R
L
have been found to be between 0 and 1
which again suggest favourable adsorption. FTIR analysis showed
that the main functional sites taking part in the sorption of CV
included carboxyl and hydroxyl groups. GFP could be regenerated
and could be reused. Column studies have shown that GFP can be
used efciently for continuous removal of CV. The present study
concludes that the GFP could be employed as low-cost adsorbent
as an alternative to commercial activated carbon for the removal of
colour and dyes fromwater and wastewater, in general and for the
removal of CV in particular.
Acknowledgements
Asma Saeed acknowledges project grant no. W/3781-2 by Inter-
national Foundation for Science, Stockholm, Sweden awarded in
collaboration with COMSTECH, Islamabad, Pakistan, and Mehwish
Sharif acknowledges nancial support formHigher EducationCom-
mission (HEC), Islamabad, Pakistan, project NRPU-707 granted to
Dr. Saeed Iqbal Zafar, HEC Eminent Researcher, to whom both are
further thankful for his guidance throughout the course of the
present study.
References
[1] A. Bhatnagar, A.K. Jain, M.K. Mukul, Removal of congo red dye fromwater using
carbon slurry waste, Environ. Chem. Lett. 2 (2005) 199202.
[2] Z. Aksu, Application of biosorption for the removal of organic pollutants: a
review, Process Biochem. 40 (2005) 9971026.
[3] C.K. Lee, K.S. Low, P.Y. Gan, Removal of some organic dyes by acid treat spent
bleaching earth, Environ. Technol. 20 (1999) 99104.
[4] M. Ramakrishnan, S. Nagarajan, Utilization of waste biomass for the removal
of basic dye from water, World Appl. Sci. J. 5 (2009) 114121.
[5] S.D. Khattri, M.K. Singh, Colour removal from synthetic dye wastewater using
a biosorbent, Water Air Soil Pollut. 120 (2000) 283294.
[6] P. Monash, G. Pugazhenthi, Adsorption of crystal violet dye fromaqueous solu-
tionusing mesoporous materials synthesizedat roomtemperature, Adsorption
15 (2009) 390405.
[7] H. Ali, S.K. Muhammad, Biosorption of crystal violet fromwater on leaf biomass
of Calotropis procera, J. Environ. Sci. Technol. 1 (2008) 143150.
[8] Y. Fu, T. Viraraghavan, Removal of congo red from aqueous solution by fungus
Aspergillus niger, Adv. Environ. Res. 7 (2002) 239247.
[9] F. Atmani, A. Bensmaili, N.Y. Mezenner, Synthetic textile efuent removal by
skin almonds waste, J. Environ. Sci. Technol. 2 (2009) 53169.
[10] E. Lorenc-Grabowska, E. Gryglewicz, Adsorption characteristics of congo
red on coal based mesoporous activated carbon, Dyes Pigments 74 (2007)
3440.
[11] C.C. Chen, H.J. Liao, C.Y. Cheng, C.Y. Yen, Y.C. Chung, Biodegradation of crystal
violet by Pseudomonas putida, Biotechnol. Lett. 29 (2007) 391396.
[12] O. Gezici, M. Kckosmano glu, A. Ayar, The adsorptionbehaviour of crystal vio-
let in functionalized sporopollenin-mediated column arrangements, J. Colloid
Interface. Sci. 304 (2006) 307316.
[13] F. Akbal, Adsorption of basic dyes fromaqueous solution onto pumice powder,
J. Colloid Interface Sci. 286 (2005) 455458.
[14] S. Senthilkumaar, P. Kalaamani, C.V. Subburaam, Liquid phase adsorption of
crystal violet onto activated carbons derived frommale owers of coconut tree,
J. Hazard. Mater. 136 (2006) 800808.
[15] V. Meshko, L. Markovska, M. Mincheva, A.E. Rodrigues, Adsorption of basic
dyes on granular activated carbon and natural zeolite, Water Res. 33 (2001)
33573366.
[16] K. Mohantay, J.T. Naidu, B.C. Meikap, M.N. Biswas, Removal of crystal violet from
wastewater by activated carbons prepared fromrice husk, Ind. Eng. Chem. Res.
45 (2006) 51655171.
[17] USDA/NASS/FFO, Citrus Summary 200607. United States Department of
Agriculture, National Agricultural Statistics Service and Florida Field Ofce,
Orlando, Florida, USA, 2008, p. 55.
[18] M.R. Wilkins, W. Widmer, K. Grohmann, R.G. Cameron, Hydrolysis of grapefruit
peel waste withcellulase andpectinase enzymes, Bioresour. Technol. 98(2007)
15961601.
[19] S.V. Ting, E.J. Deszyck, The carbohydrates in the peel of oranges and grapefruit,
J. Food Sci. 26 (1961) 146152.
[20] R.S. Blackbum, Natural Polysaccharides, Their interactions with dye molecules:
applications in efuent treatment, Environ. Sci. Technol. 38 (2004) 4905
4909.
[21] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica, and plat-
inum, J. Am. Chem. Soc. 40 (1918) 13611403.
[22] H.M.F. Freundlich, ber die adsorption in lsungen, Z. Physikalische Chem.
(Leipzig) 57A (1906) 385470.
[23] S. Lagergren, Zur theorie der sogenannten adsorption gelsterstoffe, Kungliga
Svenska Vetenskapsakademiens, Handlingar 24 (1898) 139.
[24] Y.S. Ho, G. Mackay, Sorption of dye from aqueous solution by peat, Chem. Eng.
J. 70 (1998) 115124.
[25] Y.S. Ho, Effect of pH on lead removal from water using tree fern as the sorbent,
Bioresour. Technol. 96 (2005) 12921296.
[26] S.I. Zafar, M. Bisma, A. Saeed, M. Iqbal, FTIR spectrophotometry, kinetics and
adsorption isotherms modeling, and SEM-EDX analysis for describing mech-
anism of biosorption of the cationic basic dye methylene blue by a new
biosorbent (sawdust of silver r; Abies pindrow), Fresen. Environ. Bull. 17(2008)
21092121.
[27] A. Saeed, M. Iqbal, S.I. Zafar, Immobilization of Trichoderma viride for enhanced
methylene blue biosorption: batch and column studies, J. Hazard. Mater. 168
(2009) 406415.
[28] M. Iqbal, A. Saeed, Biosorption of reactive dye by loofa sponge-immobilized
fungal biomass of Phanerochaete chrysosporium, Process Biochem. 42 (2007)
11601164.
[29] M. Iqbal, A. Saeed, S.I. Zafar, FTIR spectrophotometry, kinetics and adsorption
isotherms modeling, ion exchange, and EDX analysis for understanding the
mechanism of Cd
2+
and Pb
2+
removal by mango peel waste, J. Hazard. Mater.
164 (2009) 161171.
[30] F. Atmani, A. Bensmaili, N.Y. Mezenner, Synthetic textile efuent removal by
skin almond waste, J. Environ. Sci. Technol. 2 (2009) 153169.
[31] H.C. Chu, K.M. Chen, Reuse of activatedsludge biomass: I. Removal of basic dyes
from wastewater by biomass, Process Biochem. 37 (2002) 595600.
[32] K.R. Hall, L.C. Eagleton, A. Acrivos, T. Vermeulen, Pore and solid diffusion kinet-
ics in xed bed adsorption under constant-pattern conditions, Ind. Eng. Chem.
Fundam. 5 (1996) 212223.
[33] K.K. Krishnani, M.X. Xiaoguang, C. Christodoulatos, V.M. Boddu, Biosorption
mechanism of nine different heavy metals onto biomatrix from rice husk, J.
Hazard. Mater. 153 (2008) 12221234.
572 A. Saeed et al. / Journal of Hazardous Materials 179 (2010) 564572
[34] R. Gnanasambandam, A. Protor, Determinationof pectindegreeof esterication
by diffuse reectance Fourier transform infrared spectroscopy, Food Chem. 68
(2000) 327332.
[35] F.T. Li, H. Yang, Y. Zhao, R. Xu, Novel modiedpectinfor heavymetal adsorption,
Chin. Chem. Lett. 18 (2007) 325328.
[36] N.V. Farinella, G.D. Matos, M.A.Z. Arruda, Grape bagasse as a potential biosor-
bent of metals in efuent treatment, Bioresour. Technol. 98 (2007) 19401946.
[37] G. Guibaud, N. Tixier, A. Bouju, M. Baudu, Relationship between extracellular
polymers composition and its ability to complex Cd, Cu and Ni, Chemosphere
52 (2003) 17011710.
[38] A. Adak, M. Bandyopadhyay, A. Pal, Removal of crystal violet dye from
wastewater by surfactant-modied alumina, Sep. Purif. Technol. 44 (2005)
139144.

Potrebbero piacerti anche